Rough-strip energy dissipators (R-SEDs) can be arranged at the bend bottom of curved spillways to dissipate energy and divert flow for bend flow. Using the entropy weight and TOPSIS methods, a multi-criteria evaluation system was established for comprehensive energy dissipation and flow diversion effects of R-SEDs. Orthogonal tests and numerical simulation were conducted to analyze factors affecting these effects (average R-SED height, R-SED angle, R-SED spacing, bend width, bend centerline radius and discharge flow rate). It was found that bend width and bend centerline radius significantly affected R-SEDs' energy dissipation effects. Average R-SED height, R-SED spacing and bend centerline radius significantly affected R-SEDs' flow diversion effects. Bend width, average R-SED height and bend centerline radius significantly affected R-SEDs' combined effects of energy dissipation and flow diversion. Their energy dissipation effects were larger than the flow diversion effects. R-SEDs can effectively alleviate adverse hydraulic phenomena in curved spillways. With the recommended parameters, R-SEDs showed the best performance, with the energy dissipation rate increasing by 18.67% and the water surface superelevation coefficient decreasing by 26.14%. The accuracy of the multi-criteria evaluation system was verified. This study can provide a reference for the R-SED design of similar curved spillways.

  • R-SEDs can effectively alleviate adverse hydraulic phenomena in curved spillways.

  • A multi-criteria evaluation system was established to assess R-SEDs’ performance.

  • Factors affecting energy dissipation and flow diversion of R-SEDs were analyzed.

  • Energy dissipation rate and water surface superelevation coefficient were constructed.

  • Optimal parameters were recommended for the best performance of R-SEDs.

h

Average R-SED height

α

R-SED angle

s

R-SED spacing

w

Bend width

R

Bend centerline radius

Q

Discharge flow rate

i

Slope along the course

L1

Length of the straight inlet section

L2

Length of the straight outlet section

h1

R-SED height at the concave bank

h2

R-SED height at the convex bank

O

The center of the curvature of the bend

δ

R-SED thickness

H

Water depth at the measurement point

H0

Head over the weir

C0

Discharge coefficient of the weir

P1

Height of the weir

w1

Width of the flow diversion canal at the upstream weir

η

Energy dissipation rate

η(i)

Energy dissipation rate under the ith condition

TE1

Total mechanical energy per unit weight of water in the upstream flow cross-section

TE2

Total mechanical energy per unit weight of water in the downstream flow cross-section

SS

Water surface superelevation coefficient

SS(i)

Water surface superelevation coefficient under the ith condition

σi

Standard deviation of the superelevation of transverse water surface (Δφ) of all calculated sections under the ith condition

μi

Mean value of superelevation of transverse water surface of all calculated sections under the ith condition

Δφij

Difference between the water surface at the concave bank of the bend and the horizontal plane across the center at the jth calculated section under the ith condition

wi

Water surface width of the open channel based on the centerline water surface elevation under the ith condition

Δφ

Transverse water surface superelevation of the bend

Wj

Entropy weight

pij

Proportion of the ith evaluation object under the jth index

Ej

Entropy value of the jth index

,

Euclidean distance and between each object and the ideal solution

C

Relative closeness

P

Significance factor

DF

Degree of freedom

Adj SS

Adjusted sum of squared deviation from mean

Adj MS

Adjusted mean squared error

F

Chi-square test factor

η*

Membership degree of energy dissipation (η)

SS*

Membership degree of water surface superelevation coefficient (SS)

k

Turbulent kinetic energy

ε

Turbulent kinetic energy dissipation rate

αw

Volume fraction of water in the calculation area with respect to the calculation element

αα

Volume fraction of air in the calculation area with respect to the calculation element

T

Turbulent kinetic energy

The spillway is a key component of reservoir discharge structures (Kells & Smith 1991; Zhang et al. 2015; Damarnegara et al. 2020), mainly employed for energy dissipation and flood control. Influenced by topographic and geological conditions, construction conditions and engineering economy, some spillways must have corners, thus forming bends at the corner (Seo & Shin 2018; Yang et al. 2019). The water flowing through the bend is called bend flow, which is different from straight-section flow. When water flows through a bend, centrifugal inertia forces are generated by the curved movement of the bend flow. Thus, the surface and bottom flow migrate to the concave and convex banks, respectively, forming a closed transverse circulation within the bend section (Seyedashraf & Akhtari 2015). This circulation is combined with the longitudinal flow, forming a spiral flow. This spiral flow moves forward in the mainstream direction (Johannesson & Parker 1989). In addition, the water depth increases at the concave bank of the bend while decreasing at the convex bank. This causes a transverse gradient of the water surface (Dietrich et al. 1979).

Since Thomson (1876) revealed the existence of both longitudinal and lateral motion in bend flow, studies on bend flow have focused on two aspects, i.e., hydraulic characteristics and engineering measures. The studies on the hydraulic characteristics of bend flow mainly focus on the water depth distribution (Bathurst & Hey 1979; Molls & Chaudhry 1995; Qin et al. 2016; Zhou et al. 2017; Maatooq & Hameed 2020), flow velocity distribution (De Vriend & Geldof 1983; Anwar 1986; Odgaard & Bergs 1988; Ye & McCorquodale 1998; Han et al. 2011; Vaghefi et al. 2015; Moncho-Esteve et al. 2018; Pradhan et al. 2018; Schreiner et al. 2018; Hu et al. 2019; Kim et al. 2020; Yan et al. 2020) and secondary flow evolution (Jin & Steffler 1993; Booij 2003; Huai et al. 2012; Ramamurthy et al. 2013; Engel & Rhoads 2016; Gu et al. 2016; Shaheed et al. 2021). These studies generally describe the characteristics and evolution of bend flow and provide a theoretical basis for further studies. Based on the hydraulic characteristics of bend flow, some engineering measures in the bend have been proposed using physical model tests or numerical simulation in order to improve the bend flow structure. For example, Zhang et al. (2015) arranged continuous guide walls at the central axis of the bend to improve the bend flow pattern. However, the height of the guide wall needs to be determined based on the water depth at the bend inlet. The guide wall only applies to the bend with a small centerline radius. Thus, the practicality needs to be improved. Yang et al. (2019) arranged permeable spur dikes at the concave bank of the bend to improve the flow pattern of ‘backwater at the concave bank’. The locations with permeable spur dikes are significantly prone to congestion, which is not conducive to the structural stability of permeable spur dikes during long-term operation. Ranjan et al. (2006) arranged vanes at the central axis of the bend to reduce the secondary flow intensity of the bend flow. However, the vanes do not have energy dissipation effects and cannot dissipate the excess energy of the bend flow. Martin-Vide et al. (2010) arranged ripraps at the concave bank of the bend to reduce the scouring of bend flow at the concave bank. However, this did not change the flow pattern of ‘increasing water depth at the concave bank and decreasing water depth at the convex bank’ in the bend. In this paper, the rough-strip energy dissipators (R-SEDs) were arranged at the bend bottom of the curved spillway. These R-SEDs are simple in shape and convenient in construction and have achieved better performance in improving the flow structure and energy dissipation in the bend (Zhang et al. 2022a, 2022b).

The study on R-SEDs was mainly based on the hydraulic model test of the curved spillway of the Yin'ejike 635 Reservoir in Xinjiang, China (the geometric scale was 1:50). The physical model test was completed in the curved spillway flume of the Xinjiang Key Laboratory of Water Conservancy Engineering Safety and Water Disaster Prevention, China. At a discharge flow rate of 800 m3/s, the water depth and flow velocity at the concave and convex banks were significantly different and the bend flow was turbulent. To tackle these adverse flow patterns, R-SEDs have been arranged at the bend bottom and have shown high effectiveness in energy dissipation and flow diversion (Zhang et al. 2022b). Since then, the R-SEDs have been increasingly investigated. For example, Zhang et al. (2022a) mainly analyzed the results of the R-SEDs in the hydraulic model test of the curved spillway of the Yin'ejike 635 Reservoir using the single-factor test method. They only considered the effects of R-SED arrangement parameters on the energy dissipation and flow diversion effects of R-SEDs. The effects of engineering parameters of the curved spillway on these effects were not included. Zhang et al. (2022b) derived an equation for calculating the energy dissipation rate of R-SEDs through physical model tests, without considering the flow diversion effect of R-SEDs on bend flow. Thus, these studies still have some limitations.

Therefore, in this paper, the R-SEDs were arranged at the bend bottom of the curved spillway. Orthogonal tests and numerical simulation were performed to evaluate the R-SED arrangement parameters (average R-SED height h, R-SED angle α, R-SED spacing s) and the spillway engineering parameters (bend width w, bend centerline radius R, discharge flow rate Q) that affect the energy dissipation and diversion effects of R-SEDs. The energy dissipation rate, water surface superelevation coefficient and relative closeness were selected as response variables. Based on the entropy weight and technique for order performance by similarity to ideal solution (TOPSIS) methods, a multi-criteria evaluation system was established to analyze these parameters and determine their optimal values. Thus, an optimal parameter combination was recommended, i.e., h = 1.8 cm, s = 24 cm, α = 18°, R = 200 cm, w = 80 cm and Q = 20.0 L/s. This can provide a reference for the R-SED design of similar curved spillways. In addition, the established water surface superelevation coefficient calculation model is simple and easy to understand, facilitating its application in engineering design.

Orthogonal test design

Test apparatus

The test was conducted in a curved spillway flume at the Xinjiang Key Laboratory of Hydraulic Engineering Safety and Water Disaster Prevention, China. The test apparatus was designed according to Chinese standards, i.e., Specification for Normal Hydraulic Model Test (SL 155-2012) (Ministry of Water Resources of China 2012) and Test Regulation for Special Hydraulic Model (SL 156 ∼ 165-95) (Ministry of Water Resources of China 1995). The testing system includes a model test section and a water circulation system. The overall layout is shown in Figure 1. The model test section is divided into a straight inlet section, a bend section and a straight outlet section along the discharge direction of the spillway. The slope along the course (i = 0.025) and the straight inlet section (L1 = 60 cm) can ensure the smooth flow of inlet water into the bend section. According to Table 2, nine bends of different sizes (i.e., nine parameter combinations (R × w), including 140 cm × 50 cm, 170 cm × 60 cm, 200 cm × 80 cm, 140 cm × 80 cm, 170 cm × 50 cm, 200 cm × 60 cm, 200 cm × 50 cm, 140 cm × 60 and 170 cm × 80 cm) were designed and prepared for the test, with a bend angle of 60°. The straight outlet section (L2 = 140 cm) can smoothly connect with the flow out of the bend. The water circulation system consists of a water pump, a rectangular water storage tank, a water measuring weir, a water storage reservoir and water flow diversion pipelines.
Figure 1

Schematic diagram of test apparatus: (a) test apparatus arrangement (top view), (b) three-dimensional physical model structure of the curved spillway.

Figure 1

Schematic diagram of test apparatus: (a) test apparatus arrangement (top view), (b) three-dimensional physical model structure of the curved spillway.

Close modal

Test program

Orthogonal experimental design, referred to as orthogonal design, is a method to scientifically arrange and analyze multi-factor tests using orthogonal tables. It is one of the most commonly used experimental design methods. In the orthogonal design, representative points that meet the requirements of balance and orthogonality can be selected from the whole test and tested so as to obtain the overall stability of the development trend (Wang et al. 2022; Yuan et al. 2022).

Considering the influence of R-SED arrangement parameters and spillway engineering parameters on the energy dissipation and flow diversion effect of R-SEDs, six influencing factors were selected, including average R-SED height (h), R-SED spacing (s), R-SED angle (α), bend centerline radius (R), bend width (w) and discharge flow rate (Q). Based on the previous findings of installing R-SEDs (Zhang et al. 2022a, 2022b), three levels were selected for each factor: (−1), 0 and (+1). The orthogonal table L18(37) was selected for the test design. The factors and levels of the orthogonal test are shown in Table 1. In addition, the orthogonal test does not consider the effect of interlevel interactions. Thus, the evaluation index is only affected by the factors' independent effects (i.e., main effects) and random errors (i.e., residuals).

Table 1

Factors and levels of the orthogonal test

Levelh (cm)s (cm)α (°)R (cm)w (cm)Q (L/s)
−1 1.2 18 18 140 50 20.0 
1.5 24 22 170 60 22.5 
+1 1.8 30 26 200 80 25.0 
Levelh (cm)s (cm)α (°)R (cm)w (cm)Q (L/s)
−1 1.2 18 18 140 50 20.0 
1.5 24 22 170 60 22.5 
+1 1.8 30 26 200 80 25.0 
Table 2

Orthogonal test scheme

Runh (cm)s (cm)α (°)R (cm)w (cm)Q (L/s)
1.2 18 18 140 50 20.0 
1.2 24 22 170 60 22.5 
1.2 30 26 200 80 25.0 
1.5 18 18 170 60 25.0 
1.5 24 22 200 80 20.0 
1.5 30 26 140 50 22.5 
1.8 18 22 140 80 22.5 
1.8 24 26 170 50 25.0 
1.8 30 18 200 60 20.0 
10 1.2 18 26 200 60 22.5 
11 1.2 24 18 140 80 25.0 
12 1.2 30 22 170 50 20.0 
13 1.5 18 22 200 50 25.0 
14 1.5 24 26 140 60 20.0 
15 1.5 30 18 170 80 22.5 
16 1.8 18 26 170 80 20.0 
17 1.8 24 18 200 50 22.5 
18 1.8 30 22 140 60 25.0 
Runh (cm)s (cm)α (°)R (cm)w (cm)Q (L/s)
1.2 18 18 140 50 20.0 
1.2 24 22 170 60 22.5 
1.2 30 26 200 80 25.0 
1.5 18 18 170 60 25.0 
1.5 24 22 200 80 20.0 
1.5 30 26 140 50 22.5 
1.8 18 22 140 80 22.5 
1.8 24 26 170 50 25.0 
1.8 30 18 200 60 20.0 
10 1.2 18 26 200 60 22.5 
11 1.2 24 18 140 80 25.0 
12 1.2 30 22 170 50 20.0 
13 1.5 18 22 200 50 25.0 
14 1.5 24 26 140 60 20.0 
15 1.5 30 18 170 80 22.5 
16 1.8 18 26 170 80 20.0 
17 1.8 24 18 200 50 22.5 
18 1.8 30 22 140 60 25.0 

The R-SEDs were arranged in the bend of the spillway. Each R-SED extended continuously from the concave bank to the convex bank, close to the bend bottom. Without R-SEDs, the bend flow showed increasing water depth at the concave bank and decreasing water depth at the convex bank. In order to improve adverse bend flow, R-SED height at the concave bank (h1) was designed to be larger than the height at the convex bank (h2), i.e., h1 > h2. Therefore, R-SEDs had a trapezoidal longitudinal section. Average R-SED height (h) is the average value of the R-SED height at the concave bank (h1) and the height at the concave bank (h2), i.e., h= (h1 + h2)/2. R-SED spacing (s) is the straight line distance between the center of two adjacent R-SEDs. R-SED angle (α) is the angle between the R-SED centerline and the direction perpendicular to the bend centerline. Bend centerline radius (R) is the distance between the bend centerline and the center of the curvature of the bend (O). Bend width (w) is the horizontal distance between the two banks of the spillway. Discharge flow rate (Q) is the discharge flow rate of the spillway during regular operation. The schematic diagram of each parameter is shown in Figure 2. The orthogonal tests were designed according to Table 1. The detailed test program is shown in Table 2.
Figure 2

Schematic diagram of each factor in the orthogonal test and the cross-sectional and longitudinal structure of R-SEDs.

Figure 2

Schematic diagram of each factor in the orthogonal test and the cross-sectional and longitudinal structure of R-SEDs.

Close modal

Measurement arrangement

Test measurement instruments were selected and arranged according to the Chinese standard ‘Calibration Method of Common Instruments for Hydraulic and River Model Test’ (SL 233-2016) (Ministry of Water Resources of China 2016).

  • (a)

    Water depth measurement

The water level measurement probe was used to measure the water depth with an accuracy of 0.1 mm. A total of 51 water depth measurement cross-sections were arranged along the spillway model, i.e., #0 ∼ #50. Each cross-section was arranged with 11 measurement points (i.e., A ∼ K). To reduce the disturbance of the sidewall to the flow, two near-bank measurement points (A and E) were located 1 cm from the sidewalls of the concave and convex banks, respectively.

  • (b)

    Flow velocity measurement

The flow velocity was measured using a Pitot tube, with an accuracy of 0.1 mm. A total of 13 measurement cross-sections (i.e., #0, #4, #8, #12, #16, #20, #24, #28, #32, #36, #40, #44 and #48) were selected as flow velocity measurement cross-sections. Six measurement points (i.e., A, C, E, G, I and K) were selected for each cross-section. The vertical measurement position of each measurement point was located at 2H/3 from the bottom (H is the water depth at the measurement point).

The measurement cross-sections of water depth and flow velocity and the location of measurement points are shown in Figure 3.
  • (c)

    Flow rate measurement

Figure 3

Schematic diagram of the model measurement cross-section and measurement point arrangement.

Figure 3

Schematic diagram of the model measurement cross-section and measurement point arrangement.

Close modal
A right triangular thin-walled weir was used to measure the discharge flow rate from the spillway. The form of the weir is shown in Figure 4. The discharge flow rate is expressed as
(1)
where Q is the discharge flow rate (L/s); H0 is the head over the weir (m); C0 is the discharge coefficient of the weir, which is related to the size of the opening and can be calculated by
(2)
where P1 is the height of the weir (m); w1 is the width of the flow diversion canal at the upstream weir (m).
Figure 4

Schematic illustration of the opening shape of the right triangular thin-walled weir.

Figure 4

Schematic illustration of the opening shape of the right triangular thin-walled weir.

Close modal

Construction of evaluation indexes

Energy dissipation rate

The R-SEDs have a dual effect of energy dissipation and flow diversion on the bend flow. Evaluation indexes need to be constructed to assess the energy dissipation effect and flow diversion effect of the R-SEDs in the 18 orthogonal tests.

In order to quantify the energy dissipation effect of the continuous energy dissipation process of the R-SEDs in the bend, the energy dissipation rate (η) was introduced as an evaluation index. The change interval of the energy dissipation rate is (0, 1), and a larger energy dissipation rate indicates a larger energy dissipation effect. The energy dissipation rate is calculated through
(3)
(4)
(5)
where η(i) is the energy dissipation rate under the ith condition (%); TE1 is the total mechanical energy per unit weight of water in the upstream flow cross-section (m); Z1 is the minimum elevation of the upstream flow cross-section (m); H1 is the average water depth of the upstream flow cross-section (m); v1 is the average velocity of the upstream flow cross-section (m/s); TE2 is the total mechanical energy per unit weight of water in the downstream flow cross-section (m); Z2 is the minimum elevation of the downstream flow cross-section (m); H2 is the average water depth of the downstream flow cross-section (m); v2 is the average velocity of the downstream flow cross-section (m/s); α0 is the kinetic energy correction coefficient (. α0 depends on the flow velocity distribution at the flow cross-section. For the gradually varied flow, α0 = 1.0–1.05 and is commonly taken as 1.0 in engineering practice. In this paper, the bend flow in the spillway belonged to a nonuniform gradually-varied flow. Thus, α0 was taken as 1.0) (Zhang et al. 2022a); g is the acceleration of gravity, taken as 9.81 m/s2.

Cross-sections #8 and #32 were selected as the upstream and downstream sections of the bend, respectively. The horizontal plane where the bottom elevation of Cross-section #32 was located was taken as the reference plane. The energy dissipation rate in the 18 test scenarios was calculated using Equations (3)–(5). The calculation results of the energy dissipation rate are shown in Table 3.

Table 3

Orthogonal test results

Runh (cm)s (cm)α (°)R (°)w (cm)Q (cm)η (–)SS (–)
1.2 18 18 140 50 20.0 0.305 0.466 
1.2 24 22 170 60 22.5 0.407 0.359 
1.2 30 26 200 80 25.0 0.415 0.241 
1.5 18 18 170 60 25.0 0.358 0.488 
1.5 24 22 200 80 20.0 0.498 0.386 
1.5 30 26 140 50 22.5 0.278 0.414 
1.8 18 22 140 80 22.5 0.354 0.555 
1.8 24 26 170 50 25.0 0.322 0.456 
1.8 30 18 200 60 20.0 0.429 0.489 
10 1.2 18 26 200 60 22.5 0.396 0.328 
11 1.2 24 18 140 80 25.0 0.434 0.408 
12 1.2 30 22 170 50 20.0 0.346 0.272 
13 1.5 18 22 200 50 25.0 0.332 0.384 
14 1.5 24 26 140 60 20.0 0.314 0.467 
15 1.5 30 18 170 80 22.5 0.445 0.367 
16 1.8 18 26 170 80 20.0 0.413 0.523 
17 1.8 24 18 200 50 22.5 0.360 0.438 
18 1.8 30 22 140 60 25.0 0.276 0.497 
Runh (cm)s (cm)α (°)R (°)w (cm)Q (cm)η (–)SS (–)
1.2 18 18 140 50 20.0 0.305 0.466 
1.2 24 22 170 60 22.5 0.407 0.359 
1.2 30 26 200 80 25.0 0.415 0.241 
1.5 18 18 170 60 25.0 0.358 0.488 
1.5 24 22 200 80 20.0 0.498 0.386 
1.5 30 26 140 50 22.5 0.278 0.414 
1.8 18 22 140 80 22.5 0.354 0.555 
1.8 24 26 170 50 25.0 0.322 0.456 
1.8 30 18 200 60 20.0 0.429 0.489 
10 1.2 18 26 200 60 22.5 0.396 0.328 
11 1.2 24 18 140 80 25.0 0.434 0.408 
12 1.2 30 22 170 50 20.0 0.346 0.272 
13 1.5 18 22 200 50 25.0 0.332 0.384 
14 1.5 24 26 140 60 20.0 0.314 0.467 
15 1.5 30 18 170 80 22.5 0.445 0.367 
16 1.8 18 26 170 80 20.0 0.413 0.523 
17 1.8 24 18 200 50 22.5 0.360 0.438 
18 1.8 30 22 140 60 25.0 0.276 0.497 

Water surface superelevation coefficient

In order to quantify the flow diversion effect of the R-SEDs on the bend flow, the water surface superelevation coefficient (SS) was introduced as an evaluation index. SS varied between (0, 1) and is negatively correlated with flow diversion effects, i.e., a smaller SS indicates better flow diversion effects of the R-SED. SS is expressed as
(6)
(7)
(8)
(9)
where i indicates the sequence number of the condition; j indicates the sequence number of the calculated cross-section; SS(i) is the water surface superelevation coefficient under the ith condition; σi is the standard deviation of the superelevation of transverse water surface (Δφ) of all calculated sections under the ith condition (m); μi is the mean value of superelevation of transverse water surface of all calculated sections under the ith condition (m); Δφij is the difference between the water surface at the concave bank of the bend and the horizontal plane across the center at the jth calculated cross-section under the ith condition (m); k0 is the superelevation coefficient, taken as 0.5 for a simple circular-curved bend of a rectangular open channel; vij is the average velocity of the jth calculated section under the ith condition (m/s); wi is the water surface width of the open channel based on the centerline water surface elevation under the ith condition (m); g is the acceleration of gravity, taken as 9.81 m/s2; ri is the bend centerline radius under the ith condition (m). The calculation of the transverse water surface superelevation of the bend (Δφ) is illustrated in Figure 5.
Figure 5

Schematic diagram of the calculation of the transverse water surface superelevation of the bend (Δφ).

Figure 5

Schematic diagram of the calculation of the transverse water surface superelevation of the bend (Δφ).

Close modal

To fully measure the flow diversion effect of R-SEDs, Cross-sections #8 ∼ #48 (41 cross-sections in total) were selected to calculate the water surface superelevation coefficient. The water surface superelevation coefficients were calculated using Equations (6)–(9) for 18 orthogonal tests. The calculation results are shown in Table 3.

Multi-criteria evaluation based on entropy weight method and TOPSIS method

  • (a)

    Establishing an evaluation system of the combined effect of energy dissipation and flow diversion

Eighteen groups of orthogonal tests were used as feasibility study schemes, and the energy dissipation rate and the water surface superelevation coefficient were used as target variables to construct the original matrix .

  • (b)

    Determining the weight of the energy dissipation rate and the water surface superelevation coefficient using the entropy weight method

The entropy weight method is to calculate the entropy weight of each index according to the degree of variation of each index by using the information entropy and then to calibrate the weight of each index by the entropy weight so as to obtain a more objective index weight (Cheng et al. 2021). The matrix Y was derived from the nondimensionalization of the original matrix and then the entropy weight Wj of the energy dissipation rate and the water surface superelevation coefficient was determined. A larger entropy weight indicates that the evaluation index is more important. The calculation formula is expressed as
(10)
(11)
(12)
(13)
where rij indicates the jth evaluation index of the ith evaluation objects; i denotes the number of evaluation objects (i = 1, 2, … , m); j denotes the number of evaluation indexes (j = 1, 2, … , n); m evaluation objects refer to m test runs, and each test run is an evaluation object; n evaluation indexes refer to the energy dissipation rate and the water surface superelevation coefficient; pij is the proportion of the ith evaluation object under the jth index; Ej is the entropy value of the jth index.
  • (c)

    Obtaining the overall ranking of each program using the TOPSIS method

The TOPSIS method is commonly used to evaluate the relative strengths and weaknesses of existing objects by ranking them according to their closeness to the ideal target (Wang et al. 2019). Firstly, the original matrix is normalized to obtain a matrix and the matrix B is multiplied by the entropy weight Wj to obtain the weighting matrix Z. Secondly, the Euclidean distance and between each object and the ideal solution as well as the relative closeness C of each object is calculated. A larger relative closeness indicates that the solution is closer to the ideal solution and that the rating is better. Finally, the solutions are comprehensively ranked according to the relative closeness of each solution to form a decision basis. The calculation equations are as follows:
(14)
(15)
(16)
(17)
(18)
where z* j is the optimal solution (positive ideal solution) for the jth evaluation index; z- j is the worst solution (negative ideal solution) for the jth evaluation index.

In summary, the established multi-criteria evaluation system consists of two sub-evaluation indexes (the energy dissipation rate and water surface superelevation coefficient) and one comprehensive evaluation index (relative closeness).

Model development

Modeling and meshing

Firstly, a three-dimensional (3D) spillway model was established using SOLIDWORKS 2018. The 3D model included the straight inlet section, the 60° bend and the straight outlet section, as shown in Figure 1(b). Then, the calculation area of the spillway was extracted and meshed using ANSYS ICEM CFD 19.0. Considering the complex bend structure and free water surface in the spillway, the hybrid structured meshes (tetrahedral and hexahedral elements) were used to mesh the 3D calculation area. The structured meshes have the advantages of arbitrary structure and strong adaptability. Finally, the spillway model was calculated using ANSYS FLUENT 19.0.

In this study, the water depth and flow velocity of the spillway outlet cross-section (i.e., Cross-section #50 in Figure 3) were used as monitoring indexes. The mesh density was determined by changing the mesh quantity of the spillway for Run 5 in Table 2. Firstly, five different numbers of mesh elements (7 × 105, 8 × 105, 9 × 105, 1.0 × 106 and 1.1 × 106) were set for the geometric spillway model in Run 5, respectively. Then, the spillways with the five different mesh numbers were calculated using the same meshing and simulation methods, respectively. The simulated water depth and flow velocity of the spillway with five different mesh quantities were extracted and compared with the measured values from the physical model test. The comparison results are shown in Figure 6. The differences between simulated and experimental values were the most significant at 7 × 105 mesh elements. Then, the difference decreased with the number of mesh elements increasing to 8 × 105. The test results were consistent with the simulation results at 9 × 105, 1.0 × 106 and 1.1 × 106 mesh elements. The simulation results for the water surface structure of the spillway at the three mesh element numbers (9 × 105, 1.0 × 106 and 1.1 × 106) were generally the same (Figure 6(c)). Therefore, the number of mesh elements was selected as 9 × 105 for the spillway model. The mesh distribution is shown in Figure 7.
Figure 6

Comparison of the simulation and test results at the five numbers of mesh elements: (a) water depth at the spillway outlet cross-section, (b) flow velocity at the spillway outlet cross-section, (c) water surface structure of the spillway.

Figure 6

Comparison of the simulation and test results at the five numbers of mesh elements: (a) water depth at the spillway outlet cross-section, (b) flow velocity at the spillway outlet cross-section, (c) water surface structure of the spillway.

Close modal
Figure 7

Schematic diagram of mesh distribution and boundary condition setting.

Figure 7

Schematic diagram of mesh distribution and boundary condition setting.

Close modal

Turbulence model and control equations

The renormalization group (RNG) k-ɛ turbulence model was selected as the turbulence model. Derived from a rigorous statistical technique, this model considers turbulent vortices and can accelerate the calculation of additional terms of the turbulent kinetic energy dissipation rate equation. It has successfully simulated many complex flow problems (Huai et al. 2012; Ramamurthy et al. 2013; Ghazanfari-Hashemi et al. 2019). The two equations of the RNG k-ε turbulence model are as follows:

Turbulent kinetic energy (k) equation:
(19)
Turbulent kinetic energy dissipation rate (ε) equation:
(20)
and
(21)
where k is the turbulent kinetic energy; ε is the turbulent kinetic energy dissipation rate ; μ is the hydrodynamic viscosity; μeff is the effective viscosity coefficient; μt is the turbulent viscosity; ρ is the volume fraction averaged density; αk and αε are the inverse Prandtl number for k and ε, respectively; Gk represents the generation of turbulent kinetic energy k due to the mean flow velocity gradient; Eij is the time-averaged strain rate of the fluid; C1ε, C2ε and Cμ are empirical coefficients; t is time; in each of the above equations, i = 1, 2, 3, i.e., and ; j is the summation subscript. The values of the model constants in the above equations are shown in Table 4.
Table 4

Constant values used in the RNG kε turbulent flow model

αkαεC1εC2εCμη0β
1.39 1.39 1.42 1.68 0.0845 4.377 0.012 
αkαεC1εC2εCμη0β
1.39 1.39 1.42 1.68 0.0845 4.377 0.012 

Multi-phase flow model and boundary condition setting

The discharge flow through the spillway is mainly subject to gravity. Therefore, there is a clear water-air interface. The water-air two-phase flow can be captured at the free surface using the volume-of-fluid (VOF) method. The VOF method performs better in hydrodynamic models regarding the simulation of free surface problems and has been widely used (Pazooki et al. 2020; Li et al. 2022). The principle of the VOF method is to define the functions αw(x,y,z,t) and αa(x,y,z,t) as the volume fraction of water and air in the calculation area with respect to the calculation element, respectively. The sum of the water and air volume fraction in each calculation element was equal to 1 (i.e., αw + αa = 1). The control equation of αw is shown in Equation (22). Then, the location of the water–air interface can be detected by solving the following equation.
(22)
where ui is the velocity component, xi is the coordinate component and t is the time.

In this study, numerical simulations of the spillway flow were performed using the RNG k-ε turbulent flow model and the VOF method. The control volume finite element method (CVFEM) method was used to discretize the control equations. The semi-implicit SIMPLE algorithm was used to solve the velocity-pressure field coupling. PRESTO! was selected to calculate the pressure equations. The standard wall function method was used to deal with the near-wall flow.

The boundary condition setting of the spillway model is shown in Figure 7. According to the model test results, the water-air interface was set at the inlet water depth of 0.056 m. Thus, the water and air inlets were delineated. The boundary conditions of the water and air inlets were set as the velocity inlet and the pressure inlet (1 atm). The outlet boundary condition was set as the pressure outlet (1 atm). The standard no-slip solid wall boundary condition was applied to the remaining solid boundaries. The initial inlet water depth and flow velocity were 0.056 m and 0.448 m/s, respectively.

Statistical analysis of data

Minitab 21.1 software was used to perform the analysis of variance (ANOVA) and intuitive analysis on the calculated energy dissipation rate, water surface superelevation coefficient and relative closeness for the 18 orthogonal tests.

Energy dissipation rate

Analysis of variance

Using the energy dissipation rate as the response variable, ANOVA was conducted on the six influencing factors (predictor variables). The results are shown in Table 5. The significance level was selected as 0.05. When the P-value corresponding to the main effect was less than 0.05 (P < 0.05), the original hypothesis was rejected, and the total effect of the regression was considered significant. The analysis was focused on the main effects on the evaluation indexes without considering the interactions between different factors. Adjusted R2 (R2 (adj)) = 83.98% > 80.0% (Table 5) for the influencing factors, indicating that the orthogonal tests were statistically significant. Further analysis of the test levels shows that the differences between the different groups of bend centerline radius (R) and bend width (w) were significant (PR < 0.05, Pw < 0.05). This indicates that these two factors dominated the changing trend of the energy dissipation rate. The remaining factors had a test level of P > 0.05, indicating that these factors showed homogeneity of variance and did not exhibit significant differences. This indicates that the remaining factors had insignificant effects on the energy dissipation rate. The above analysis shows that the spillway engineering parameters influenced the energy dissipation rate more than the R-SED arrangement parameters. Therefore, the influence of the spillway body shape on the energy dissipation rate of R-SEDs should be highlighted in engineering design.

Table 5

Analysis of variance for the energy dissipation rate

SourceDFAdj SSAdj MSF-valueP-valueSignificance
h 0.001851 0.000926 1.51 0.306 – 
s 0.002978 0.001489 2.43 0.183 – 
α 0.003155 0.001578 2.58 0.170 – 
R 0.019343 0.009672 15.80 0.007 
w 0.032181 0.016091 26.28 0.002 
Q 0.002392 0.001196 1.95 0.236 – 
Error 0.003061 0.000612 – – – 
Total 17 0.064963 – – – – 
  R2 = 0.9529 R2 (adj) = 0.8398   
SourceDFAdj SSAdj MSF-valueP-valueSignificance
h 0.001851 0.000926 1.51 0.306 – 
s 0.002978 0.001489 2.43 0.183 – 
α 0.003155 0.001578 2.58 0.170 – 
R 0.019343 0.009672 15.80 0.007 
w 0.032181 0.016091 26.28 0.002 
Q 0.002392 0.001196 1.95 0.236 – 
Error 0.003061 0.000612 – – – 
Total 17 0.064963 – – – – 
  R2 = 0.9529 R2 (adj) = 0.8398   

Note: DF is the degree of freedom; Adj SS is the adjusted sum of squared deviation from mean; Adj MS is the adjusted mean squared error; F is the chi-square test factor; P is the significance factor.

*P < 0.05, and the regression is significant

Intuitive analysis

Intuitive analysis (i.e., range analysis) can reflect the degree of change in the test index when the level of a certain test factor changes. The energy dissipation rate was used as the response variable. The intuitive analysis results of the six influencing factors are shown in Table 6. The six factors (Table 6) were ranked in descending order in terms of their effects on the energy dissipation rate: bend width (w) > bend centerline radius (R) > R-SED angle (α) > R-SED spacing (s) > discharge flow rate (Q) > average R-SED height (h). This shows the degree of influence of each factor on the energy dissipation rate. Thus, for the R-SED design, the effects of the significant factors (bend width w and bend centerline radius R) on the energy dissipation rate should be mainly considered. The values of these significant factors can be selected based on engineering characteristics and the main effect analysis of the energy dissipation rate (Figure 8). The remaining factors (secondary factors) can be selected based on the main effect analysis of the energy dissipation rate (Figure 8) or engineering experience.
Table 6

Intuitive analysis results of the energy dissipation rate

Levelh (cm)s (cm)α (°)R (cm)w (cm)Q (L/s)
0.3838 0.3597 0.3885 0.3268 0.3238 0.3842 
0.3708 0.3892 0.3688 0.3818 0.3633 0.3733 
0.3590 0.3648 0.3563 0.4050 0.4265 0.3562 
Delta 0.0248 0.0295 0.0322 0.0782 0.1027 0.0280 
Rank 
Levelh (cm)s (cm)α (°)R (cm)w (cm)Q (L/s)
0.3838 0.3597 0.3885 0.3268 0.3238 0.3842 
0.3708 0.3892 0.3688 0.3818 0.3633 0.3733 
0.3590 0.3648 0.3563 0.4050 0.4265 0.3562 
Delta 0.0248 0.0295 0.0322 0.0782 0.1027 0.0280 
Rank 
Figure 8

Main effect plot of the energy dissipation rate (η).

Figure 8

Main effect plot of the energy dissipation rate (η).

Close modal

The characteristic mean of each factor level is first obtained. Then, the main effect plot can be achieved by connecting the response mean of each factor level using a line. Whether a factor has the main effect can be determined by comparing the slope of the regression lines. A larger slope of the regression line indicates more significant main effects. The main effect plot of the energy dissipation rate (η) is shown in Figure 8. The regression lines of bend width (w) and bend centerline radius (R) were the steepest, indicating that η was the most sensitive to the change of these two factors. The two factors (w and R) were positive indicators, i.e., at a larger factor, η is higher, and R-SEDs' energy dissipation effects are larger. η was the highest (0.427) at the largest bend width (w = 80 cm) while decreasing by 31.38% with w decreasing to 50 cm. η was the highest (0.405) at the largest bend centerline radius (R = 200 cm) while decreasing by 23.85% with R decreasing to 140 cm. The slopes of the regression lines of three factors (average R-SED height h, R-SED angle α and discharge flow rate Q) were similar, i.e., the effect of these three factors on η was consistent. These three factors were negative indicators, i.e., at a larger factor, η is lower, and R-SEDs' energy dissipation effects are lower. η was the highest (0.385) at the smallest average R-SED height (h = 1.2 cm) while decreasing by 7.54% with h increasing to 1.8 cm. η was the highest (0.388) at the smallest R-SED angle (α = 18°) while decreasing by 8.99% with α increasing to 26°. η was the highest (0.385) at the smallest discharge flow rate (Q = 20 L/s) while decreasing by 8.15% with Q increasing to 25.0 L/s. R-SED spacing (s) was a two-way indicator, i.e., s is a positive indicator within 18–24 cm and a negative indicator within 24–30 cm. η was the highest (0.389) at s = 24 cm. η increased by 8.06% with s increasing from 18 to 24 cm and increased by 6.28% with s decreasing from 30 to 24 cm.

Water surface superelevation coefficient

Analysis of variance

The water surface superelevation coefficient was used as the response variable. The significance level was selected as 0.05. The ANOVA was performed on the six influencing factors. The results are shown in Table 7. R2 (adj) = 92.08% > 80.0% for these influencing factors (Table 7), indicating that the orthogonal test was statistically significant. Further analysis shows that the test levels of average R-SED height (h), R-SED spacing (s) and bend centerline radius (R) differed significantly between groups (Ph < 0.05, Ps < 0.05 and PR < 0.05). This indicates that these three factors dominated the changing trend of the water surface superelevation coefficient. The remaining factors had a test level of P > 0.05. Thus, these factors had homogeneity of variance and did not show significant differences. Thus, they had fewer effects on the water surface superelevation coefficient. The above analysis shows that compared with the spillway engineering parameters, the R-SED arrangement parameters significantly affected the water surface superelevation coefficient.

Table 7

Analysis of variance for the water surface superelevation coefficient

SourceDFAdj SSAdj MSF-valueP-valueSignificance
h 0.065132 0.032566 57.48 0.000 
s 0.017942 0.008971 15.83 0.007 
α 0.005184 0.002592 4.58 0.074 – 
R 0.024958 0.012479 22.03 0.003 
w 0.003534 0.001767 3.12 0.132 – 
Q 0.002054 0.001027 1.81 0.256 – 
Error 0.002833 0.000567 – – – 
Total 17 0.121637 – – – – 
  R2 = 0.9767 R2 (adj) = 0.9208   
SourceDFAdj SSAdj MSF-valueP-valueSignificance
h 0.065132 0.032566 57.48 0.000 
s 0.017942 0.008971 15.83 0.007 
α 0.005184 0.002592 4.58 0.074 – 
R 0.024958 0.012479 22.03 0.003 
w 0.003534 0.001767 3.12 0.132 – 
Q 0.002054 0.001027 1.81 0.256 – 
Error 0.002833 0.000567 – – – 
Total 17 0.121637 – – – – 
  R2 = 0.9767 R2 (adj) = 0.9208   

*P < 0.05, and the regression is significant

Intuitive analysis

The water surface superelevation coefficient was used as the response variable. The intuitive analysis results of the six influencing factors are shown in Table 8. The six influencing factors were ranked in descending order in terms of their effects on the water surface superelevation coefficient: average R-SED height (h) > bend centerline radius (R) > R-SED spacing (s) > R-SED angle (α) > bend width (w) > discharge flow rate (Q). This shows the degree of influence of each factor on the water surface superelevation coefficient. Thus, for the R-SED design, the effects of the significant factors (average R-SED height (h), bend centerline radius (R) and R-SED spacing (s)) on water surface superelevation coefficient should be mainly considered. The values of these significant factors can be selected based on engineering characteristics and the main effect analysis of the water surface superelevation coefficient (Figure 9). The remaining factors (secondary factors) can be selected based on the main effect analysis of SS (Figure 9) or engineering experience.
Table 8

Intuitive analysis results of the water surface superelevation coefficient

Levelh (cm)s (cm)α (°)R (cm)w (cm)Q (L/s)
0.3457 0.4573 0.4427 0.4678 0.4050 0.4338 
0.4177 0.4190 0.4088 0.4108 0.4380 0.4102 
0.4930 0.3800 0.4048 0.3777 0.4133 0.4123 
Delta 0.1473 0.0773 0.0378 0.0902 0.0330 0.0237 
Rank 
Levelh (cm)s (cm)α (°)R (cm)w (cm)Q (L/s)
0.3457 0.4573 0.4427 0.4678 0.4050 0.4338 
0.4177 0.4190 0.4088 0.4108 0.4380 0.4102 
0.4930 0.3800 0.4048 0.3777 0.4133 0.4123 
Delta 0.1473 0.0773 0.0378 0.0902 0.0330 0.0237 
Rank 
Figure 9

Main effect plot of the water surface superelevation coefficient (SS).

Figure 9

Main effect plot of the water surface superelevation coefficient (SS).

Close modal

The main effect plot of water surface superelevation coefficient (SS) is shown in Figure 9. The regression lines of average R-SED height (h), R-SED spacing (s) and bend centerline radius (R) were the steepest. This indicates that SS was the most sensitive to the changes in these three factors. Average R-SED height (h) was a negative indicator, i.e., at a larger h, SS is higher, and R-SEDs' flow diversion effects are lower. SS was the smallest (0.345) at the smallest average R-SED height (h = 1.2 cm) while increasing by 29.88% with h increasing to 1.8 cm. R-SED spacing (s), bend centerline radius (R) and R-SED angle (α) were positive indicators, i.e., at a larger value, SS is smaller, and R-SEDs' flow diversion effects are larger. SS was the smallest (0.388) at the largest R-SED spacing (s = 30 cm) while increasing by 15.47% with s decreasing to 18 cm. SS was the smallest (0.378) at the largest bend centerline radius (R = 200 cm) while increasing by 19.40% with R decreasing to 140 cm. SS was the smallest (0.406) at the largest R-SED angle (α = 26°) while increasing by 8.56% with α decreasing to 18°. Bend width (w) and discharge flow rate (Q) were two-way indicators. The bend width (w) was a negative indicator within 50–60 cm and a positive indicator within 60–80 cm. SS is the largest at w = 60 cm, with the weakest flow diversion effect. SS decreased by 7.27% with w decreasing from 60 to 50 cm while decreasing by 5.68% from 60 to 80 cm, with improved flow diversion effects. The discharge flow rate (Q) was a positive indicator within 20.0–22.5 L/s and a negative indicator within 22.5–25.0 L/s. At Q= 22.5 L/s, SS was the smallest (0.409), and the R-SEDs had the most significant flow diversion effect.

Relative closeness

Determination of the weights of each evaluation index using the entropy weight method

The weights of each evaluation index (η and SS) were calculated using Equations (10)–(13) and the calculated parameters are shown in Table 9. The weight of the energy dissipation rate (W(η) = 0.629) was larger than that of the superelevation coefficient (W(SS) = 0.371). This indicates that the energy dissipation rate carried more information than the water surface superelevation coefficient and that the energy dissipation effect of R-SEDs on the bend flow was larger than the flow diversion effect.

Table 9

Weight of each evaluation index using the entropy weight method

RunηSSη*SS*p(η*)p(SS*)E(η)E(SS)W(η)W(SS)
0.305 0.466 0.131 0.717 0.017 0.070 0.917 0.951 0.629 0.371 
0.407 0.359 0.590 0.376 0.076 0.037 
0.415 0.241 0.626 0.000 0.081 0.000 
0.358 0.488 0.369 0.787 0.048 0.077 
0.498 0.386 1.000 0.462 0.130 0.045 
0.278 0.414 0.009 0.551 0.001 0.054 
0.354 0.555 0.351 1.000 0.046 0.098 
0.322 0.456 0.207 0.685 0.027 0.067 
0.429 0.489 0.689 0.790 0.089 0.078 
10 0.396 0.328 0.541 0.277 0.070 0.027 
11 0.434 0.408 0.712 0.532 0.092 0.052 
12 0.346 0.272 0.315 0.099 0.041 0.010 
13 0.332 0.384 0.252 0.455 0.033 0.045 
14 0.314 0.467 0.171 0.720 0.022 0.071 
15 0.445 0.367 0.761 0.401 0.099 0.039 
16 0.413 0.523 0.617 0.898 0.080 0.088 
17 0.36 0.438 0.378 0.627 0.049 0.062 
18 0.276 0.497 0.000 0.815 0.000 0.080 
RunηSSη*SS*p(η*)p(SS*)E(η)E(SS)W(η)W(SS)
0.305 0.466 0.131 0.717 0.017 0.070 0.917 0.951 0.629 0.371 
0.407 0.359 0.590 0.376 0.076 0.037 
0.415 0.241 0.626 0.000 0.081 0.000 
0.358 0.488 0.369 0.787 0.048 0.077 
0.498 0.386 1.000 0.462 0.130 0.045 
0.278 0.414 0.009 0.551 0.001 0.054 
0.354 0.555 0.351 1.000 0.046 0.098 
0.322 0.456 0.207 0.685 0.027 0.067 
0.429 0.489 0.689 0.790 0.089 0.078 
10 0.396 0.328 0.541 0.277 0.070 0.027 
11 0.434 0.408 0.712 0.532 0.092 0.052 
12 0.346 0.272 0.315 0.099 0.041 0.010 
13 0.332 0.384 0.252 0.455 0.033 0.045 
14 0.314 0.467 0.171 0.720 0.022 0.071 
15 0.445 0.367 0.761 0.401 0.099 0.039 
16 0.413 0.523 0.617 0.898 0.080 0.088 
17 0.36 0.438 0.378 0.627 0.049 0.062 
18 0.276 0.497 0.000 0.815 0.000 0.080 

Note: η* is the degree of membership function of energy dissipation (η); SS* is the degree of membership function of water surface superelevation coefficient (SS). The degree of membership is the ratio of (the index value − the minimum index value) to (the maximum index value − the minimum index value).

Determination of the overall ranking of each program using the TOPSIS method

The relative closeness can comprehensively measure the energy dissipation and flow diversion effects of R-SEDs. The relative closeness of each run to the ideal solution was calculated using Equations (14)–(18). Then, the overall ranking of the energy dissipation and flow diversion effect of R-SEDs on the bend flow at each run was obtained (Table 10). Run 5 (h = 1.5 cm, s = 24 cm, α = 22°, R = 200 cm, w = 80 cm and Q = 20L/s) had the largest relative closeness and the highest overall rating.

Table 10

Overall ranking of each run using the TOPSIS method

RunηSS1 − SSb(η)b(1−SS)z(η)z(1−SS)CRank
0.305 0.466 0.534 0.191 0.214 0.120 0.080 0.083 0.032 0.279 16 
0.407 0.359 0.641 0.255 0.198 0.160 0.074 0.054 0.057 0.514 
0.415 0.241 0.759 0.260 0.133 0.164 0.049 0.072 0.055 0.432 11 
0.358 0.488 0.512 0.224 0.270 0.141 0.100 0.057 0.060 0.513 
0.498 0.386 0.614 0.312 0.213 0.196 0.079 0.035 0.092 0.728 
0.278 0.414 0.586 0.174 0.229 0.110 0.085 0.091 0.035 0.279 16 
0.354 0.555 0.445 0.222 0.307 0.140 0.114 0.057 0.071 0.557 
0.322 0.456 0.544 0.202 0.252 0.127 0.093 0.072 0.048 0.397 12 
0.429 0.489 0.511 0.269 0.270 0.169 0.100 0.030 0.079 0.722 
10 0.396 0.328 0.672 0.248 0.181 0.156 0.067 0.061 0.051 0.451 10 
11 0.434 0.408 0.592 0.272 0.225 0.171 0.084 0.039 0.071 0.644 
12 0.346 0.272 0.728 0.217 0.150 0.136 0.056 0.083 0.028 0.254 18 
13 0.332 0.384 0.616 0.208 0.212 0.131 0.079 0.074 0.037 0.331 15 
14 0.314 0.467 0.533 0.197 0.258 0.124 0.096 0.075 0.049 0.394 13 
15 0.445 0.367 0.633 0.279 0.203 0.175 0.075 0.044 0.071 0.620 
16 0.413 0.523 0.477 0.259 0.289 0.163 0.107 0.034 0.079 0.698 
17 0.36 0.438 0.562 0.226 0.242 0.142 0.090 0.059 0.052 0.468 
18 0.276 0.497 0.503 0.173 0.274 0.109 0.102 0.088 0.052 0.372 14 
RunηSS1 − SSb(η)b(1−SS)z(η)z(1−SS)CRank
0.305 0.466 0.534 0.191 0.214 0.120 0.080 0.083 0.032 0.279 16 
0.407 0.359 0.641 0.255 0.198 0.160 0.074 0.054 0.057 0.514 
0.415 0.241 0.759 0.260 0.133 0.164 0.049 0.072 0.055 0.432 11 
0.358 0.488 0.512 0.224 0.270 0.141 0.100 0.057 0.060 0.513 
0.498 0.386 0.614 0.312 0.213 0.196 0.079 0.035 0.092 0.728 
0.278 0.414 0.586 0.174 0.229 0.110 0.085 0.091 0.035 0.279 16 
0.354 0.555 0.445 0.222 0.307 0.140 0.114 0.057 0.071 0.557 
0.322 0.456 0.544 0.202 0.252 0.127 0.093 0.072 0.048 0.397 12 
0.429 0.489 0.511 0.269 0.270 0.169 0.100 0.030 0.079 0.722 
10 0.396 0.328 0.672 0.248 0.181 0.156 0.067 0.061 0.051 0.451 10 
11 0.434 0.408 0.592 0.272 0.225 0.171 0.084 0.039 0.071 0.644 
12 0.346 0.272 0.728 0.217 0.150 0.136 0.056 0.083 0.028 0.254 18 
13 0.332 0.384 0.616 0.208 0.212 0.131 0.079 0.074 0.037 0.331 15 
14 0.314 0.467 0.533 0.197 0.258 0.124 0.096 0.075 0.049 0.394 13 
15 0.445 0.367 0.633 0.279 0.203 0.175 0.075 0.044 0.071 0.620 
16 0.413 0.523 0.477 0.259 0.289 0.163 0.107 0.034 0.079 0.698 
17 0.36 0.438 0.562 0.226 0.242 0.142 0.090 0.059 0.052 0.468 
18 0.276 0.497 0.503 0.173 0.274 0.109 0.102 0.088 0.052 0.372 14 

Analysis of variance

Relative closeness was used as the response variable. A significance level of 0.05 was selected. The ANOVA was conducted to examine the importance of the six influencing factors on the relative closeness, as shown in Table 11. R2 (adj) = 87.92% > 80.0% (Table 11) for the influencing factors, indicating that the orthogonal test was statistically significant. Further analysis shows that the differences in the test levels between groups were significant for bend width (w), average R-SED height (h) and bend centerline radius (R) (Pw < 0.05, Ph < 0.05 and PR < 0.05). This indicates that these three factors dominated the changing trend of relative closeness. The remaining factors had a test level of P > 0.05. Thus, these factors showed homogeneity of variance and did not exhibit significant differences. Thus, the remaining factors had insignificant effects on the relative closeness.

Table 11

Analysis of variance of relative closeness

SourceDFAdj SSAdj MSF-valueP-valueSignificance
h 0.03423 0.017113 2.24 0.012 
s 0.01886 0.009431 1.23 0.367 – 
α 0.03362 0.016810 2.20 0.207 – 
R 0.03382 0.016911 2.21 0.015 
w 0.23435 0.117177 15.31 0.007 
Q 0.01242 0.006211 0.81 0.495 – 
Error 0.03826 0.007653 – – – 
Total 17 0.40557 – – – – 
  R2 = 0.9057 R2 (adj) = 0.8792   
SourceDFAdj SSAdj MSF-valueP-valueSignificance
h 0.03423 0.017113 2.24 0.012 
s 0.01886 0.009431 1.23 0.367 – 
α 0.03362 0.016810 2.20 0.207 – 
R 0.03382 0.016911 2.21 0.015 
w 0.23435 0.117177 15.31 0.007 
Q 0.01242 0.006211 0.81 0.495 – 
Error 0.03826 0.007653 – – – 
Total 17 0.40557 – – – – 
  R2 = 0.9057 R2 (adj) = 0.8792   

*P < 0.05, and the regression is significant

Intuitive analysis

The relative closeness was selected as the response variable. The intuitive analysis results of the six influencing factors are shown in Table 12. The effects of the six factors on relative closeness were ranked in descending order: bend width (w) > average R-SED height (h) > bend centerline radius (R) > R-SED angle (α) > R-SED spacing (s) > discharge flow rate (Q). This shows the degree of influence of each factor on the relative closeness. Thus, for the R-SED design, the effects of the significant factors (bend width (w), average R-SED height (h), bend centerline radius (R)) on the relative closeness should be mainly considered. The values of these significant factors can be selected based on engineering characteristics and the main effect analysis of the relative closeness (Figure 10). The remaining factors (secondary factors) can be selected based on the main effect analysis of the relative closeness (Figure 10) or engineering experience.
Table 12

Intuitive analysis results of relative closeness

Levelh (cm)s (cm)α (°)R (cm)w (cm)Q (L/s)
0.4290 0.4715 0.5410 0.4208 0.3347 0.5125 
0.4775 0.5242 0.4593 0.4993 0.4943 0.4815 
0.5357 0.4465 0.4418 0.5220 0.6132 0.4482 
Delta 0.1067 0.0777 0.0992 0.1012 0.2785 0.0643 
Rank 
Levelh (cm)s (cm)α (°)R (cm)w (cm)Q (L/s)
0.4290 0.4715 0.5410 0.4208 0.3347 0.5125 
0.4775 0.5242 0.4593 0.4993 0.4943 0.4815 
0.5357 0.4465 0.4418 0.5220 0.6132 0.4482 
Delta 0.1067 0.0777 0.0992 0.1012 0.2785 0.0643 
Rank 
Figure 10

Main effect plot of relative closeness (C).

Figure 10

Main effect plot of relative closeness (C).

Close modal

The main effect plots of relative closeness (C) are shown in Figure 10. The regression lines of bend width (w), average R-SED height (h) and bend centerline radius (R) were the steepest, indicating that C was the most sensitive to the changes of these three factors. These three factors were positive indicators, i.e., at a larger factor, C is larger, and the comprehensive energy dissipation and flow diversion effects of R-SEDs are stronger. C was the largest (0.613) at the largest bend width (w = 80 cm) while decreasing by 81.90% with w decreasing to 50 cm. C was the largest (0.537) at the largest average R-SED height (h = 1.8 cm) while decreasing by 24.88% with h decreasing to 1.2 cm. C was the largest (0.524) at the largest bend centerline radius (R = 200 cm) while decreasing by 24.76% with R decreasing to 140 cm. R-SED angle (α) and discharge flow rate (Q) were negative indicators, i.e., at a larger factor, C is smaller, and the comprehensive energy dissipation and diversion effects of R-SEDs are weaker. C reached the maximum (0.542, 0.513) when these two factors were 18° and 20.0 L/s, respectively. R-SED spacing (s) was a two-way indicator: a positive indicator within 18–24 cm and a negative indicator within 24–30 cm. At s = 24 cm, C reached the maximum (0.525). Combined with the findings of Section 3.1.2, the two-way nature of R-SED spacing (s) indicates that the value of s should not be too large or too small. Excessively large R-SED spacing will reduce the number of R-SEDs, while too small spacing will reduce the contact area of water flow between adjacent R-SEDs. These are not conducive to energy dissipation and flow diversion.

The value of each factor at which the relative closeness reached its maximum was selected to obtain the combination of recommended parameters, i.e., h = 1.8 cm, s = 24 cm, α = 18°, R = 200 cm, w = 80 cm and Q = 20.0L/s.

To verify the accuracy of the multi-criteria evaluation system (i.e., to verify the advantages of the recommended parameter combination) and the necessity of arranging R-SEDs in the bend, numerical simulation was performed in three runs, including (1) Run without R-SEDs (R = 200 cm, w = 80 cm and Q = 20.0 L/s); (2) Run 5 (the optimal run in the 18 orthogonal tests: h = 1.5 cm, s = 24 cm, α = 22°, R = 200 cm, w = 80 cm and Q = 20.0 L/s); (3) the run with the recommended parameter combination (Run OPC) (h = 1.8 cm, s = 24 cm, α = 18°, R = 200 cm, w = 80 cm and Q = 20.0 L/s).

Water depth variation of the spillway

Figure 11 shows the comparison of water depth variation of the spillway in the three runs. A total of 17 typical cross-sections (Cross-sections #a ∼ #p and the outlet cross-section) were set in the spillway. The water depth variation patterns were studied comprehensively in longitudinal and transverse directions. Without R-SEDs (Figure 11(a)), the water depth of the bend (Cross-sections #c ∼ #k) was affected by centrifugal force and showed a pattern of ‘increasing at the concave bank and decreasing at the convex bank’. The water depth was generally larger at the concave bank than at the convex bank. Particularly, there was no water at the convex bank of Cross-sections #i ∼ #l. The water depth of the straight outlet section (Cross-sections #l ∼ #p and the outlet cross-section) gradually transited from ‘Concave bank > Convex bank’ to ‘Convex bank > Concave bank’. Thus, a deflected flow occurred. Figure 11(b) and 11(c) shows that the R-SED arrangement (R-SEDs #1 ∼ #9) at the bend bottom reduced the water depth difference between the two banks of the bend. This generally eliminated the deflected flow pattern that occurred in the outlet straight section without R-SEDs, indicating the flow diversion effect of the R-SEDs. It can also be found that the R-SEDs in Run 5 induced nearly the same water depth only at both banks of the outlet cross-section. However, the R-SEDs in Run OPC can help to achieve nearly the same water depth even at both banks of Cross-section #p. Therefore, the R-SEDs in Run OPC can facilitate a more rapid uniform water depth at the cross-sections.
Figure 11

Comparison of water depth variation in the spillway under three working conditions: (a) Run without R-SEDs, (b) Run 5, (c) Run OPC.

Figure 11

Comparison of water depth variation in the spillway under three working conditions: (a) Run without R-SEDs, (b) Run 5, (c) Run OPC.

Close modal

Distribution of typical cross-sectional streamlines at the bend

Figure 12 shows the superimposed distribution of streamlines and velocity contour plots for typical cross-sections of the spillway bend (i.e., Cross-sections #c ∼ #k in Figure 11(a)) in the three runs. Without R-SEDs (Figure 12(i)), the streamlines of most cross-sections were linearly distributed. At a cross-section closer to the outlet of the bend (Cross-section #k), the trend (i.e., the flow velocity at the convex bank > concave bank) was more prominent. The cross-sectional flow velocity distribution was more nonuniform. The flow velocity at the concave bank and in the middle of Cross-sections #h and #j were generally larger (up to 1.27 m/s). After the R-SEDs were arranged at the bend bottom (Figure 12(ii) and 12(iii)), the discharge flow hit the R-SEDs. This caused the flow to break and swirl. Thus, the cross-sectional flow was redistributed, showing wavy streamlines. After the water reached Cross-section #f, cross-sectional circulation gradually occurred. There were low-velocity areas (<0.5 m/s) on the upstream and downstream faces of the R-SEDs. The maximum cross-sectional flow velocity in Run OPC was reduced to 0.94 m/s by the R-SEDs and was 25.98% lower than that in Run without R-SEDs (1.27 m/s). This was due to the combined effect of centrifugal force, water pressure and R-SEDs' energy dissipation and flow diversion. Thus, the surface and bottom water flowed to the cross-section's convex and concave banks, respectively. The flow formed a clockwise cross-sectional circulation on both sides of the R-SED, effectively reducing the flow velocity. It is found that the cross-sectional circulation occurred in Cross-sections #e ∼ #h and #d ∼ #j in the bend in Run 5 and Run OPC, respectively (Figure 12(ii) and 12(iii)). This indicates that the R-SED shape under the Run OPC can extend the occurrence range of the cross-sectional circulation and facilitate solving the problem of nonuniform cross-sectional velocity distribution in the bend.
Figure 12

Superimposed distribution of streamline and velocity contour plots for typical cross-sections of the spillway bend in the three runs: Cross-sections (a) #c; (b) #d; (c) #e; (d) #f; (e) #g; (f) #h; (g) #i; (h) #j; (j) #k.

Figure 12

Superimposed distribution of streamline and velocity contour plots for typical cross-sections of the spillway bend in the three runs: Cross-sections (a) #c; (b) #d; (c) #e; (d) #f; (e) #g; (f) #h; (g) #i; (h) #j; (j) #k.

Close modal

Vorticity analysis

Vorticity is one of the most important physical quantities to describe the motion of a vortex and is defined as the curl of the fluid velocity vector. Figure 13 shows the vorticity distribution in the spillway in the three runs. The vertical distributions of vorticity at the concave bank, central axis and convex bank of Cross-section #g (the middle cross-section of the bend) were extracted. The vorticity was mainly concentrated in the spillway bend (Figure 13). Affected by the large turbulent shear stress at the solid–liquid and gas–liquid interfaces, the vorticity was larger at the position closer to the bottom and surface of the flow. The vorticity was the smallest in the middle flow. In Run without R-SEDs (Figure 13(a)), the overall vorticity in the bend was generally small. Larger vorticity was concentrated in the concave bank, central axis and convex bank of the bend. Particularly, the vorticity in the middle flow of the cross-section across the central axis was the largest (70/s). After the R-SEDs were installed at the bend bottom (Figure 13(a) and 13(b)), the mixing and swirling between the flow were enhanced. Thus, the overall vorticity increased. The swirling eddies at the concave bank, central axis and convex bank of Cross-section #g occurred in different scales. The vertical vorticity distribution at the central axis shows that the vorticity at the top corner of the R-SEDs (up to 100/s) was commonly larger due to the collision and rubbing of the flow against the top corner. The maximum vorticity (100/s) in the run with R-SEDs was 42.86% higher than the maximum vorticity (70/s) in Run without R-SEDs. From Figure 13(b) and 13(c), it is found that in Run OPC, the vorticity in the bend and the straight outlet section had a wider distribution than in Run 5. The vorticity at the concave bank, central axis and convex bank of Cross-section #g was slightly larger in Run OPC than in Run 5. This indicates that the R-SED shape under the Run OPC was more favorable to the energy dissipation in the bend.
Figure 13

Vorticity distribution in the spillway in the three runs: (a) Run without R-SEDs, (b) Run 5, (c) Run OPC.

Figure 13

Vorticity distribution in the spillway in the three runs: (a) Run without R-SEDs, (b) Run 5, (c) Run OPC.

Close modal

Turbulent kinetic energy analysis

Turbulent kinetic energy is a physical quantity that characterizes the overall turbulence of the flow and is expressed as
(23)
where T is the turbulent kinetic energy (m2/s2); u’, v’ and w’ are the pulsatile flow velocity in the longitudinal, transverse and vertical directions (m/s), respectively.
Figure 14 shows the vertical distribution of turbulent kinetic energy at the central axis of the typical cross-section of the spillway bend (i.e., Cross-sections #c ∼ #k in Figure 11(a)) in the three runs. The larger turbulent kinetic energy in each cross-section of the bend occurred below 0.5Z, and the maximum value was concentrated at the bend bottom. This is because turbulent shear stress at the solid-liquid interface was larger, resulting in larger turbulent energy. Thus, swirling eddies on different scales can be easily generated. The upward movement and mixing of the vortexes were accompanied by energy transfer and dissipation. Thus, this resulted in smaller turbulent energy in the upper flow (>0.5Z). In Run without R-SEDs, the distribution of turbulent kinetic energy in each cross-section of the bend was similar. The turbulent energy reached the maximum at the inlet of the bend (0.183 m2/s2) and then gradually decreased. After the arrangement of R-SEDs at the bend bottom (Run 5 and Run OPC), the contact area between the flows was increased by the discharge flow hitting the R-SEDs. This enhanced the overall turbulence of the flow, and the turbulent kinetic energy below 0.5Z increased. Most turbulent kinetic energy and the maximum value of each cross-section below 0.5Z in Run OPC were larger than in Run 5. The maximum value in Run OPC (0.334 m2/s2) (Figure 14(h)) increased by 82.51 and 32.02% compared to those in Run without R-SEDs (0.183 m2/s2) and Run 5 (0.253 m2/s2), respectively. In Figure 14(a)–14(d), the turbulent kinetic energy curves of Run 5 and Run OPC showed some overlaps. In Figure 14(f)–14(i), the differences between the turbulent kinetic energy of these two runs below 0.5Z became significant. This can be explained according to the findings in Section 4.2, i.e., when the water reached the middle cross-section of the bend (#g, see Figure 14(e)), the R-SEDs in Run OPC allowed the circulation to more fully developed. Under the combined effects of centrifugal force, water pressure and R-SED energy dissipation and flow diversion, the shear stress in the side wall of both spillway banks increased in Run OPC. The mixing of upper and lower flows was enhanced, resulting in a higher turbulent diffusion rate and, thus, a higher turbulence energy growth rate.
Figure 14

Vertical distribution of turbulent kinetic energy at the central axis of typical spillway bend cross-sections in the three runs: Cross-sections (a) #c, (b) #d, (c) #e, (d) #f, (e) #g, (f) #h, (g) #i, (h) #j, (i) #k.

Figure 14

Vertical distribution of turbulent kinetic energy at the central axis of typical spillway bend cross-sections in the three runs: Cross-sections (a) #c, (b) #d, (c) #e, (d) #f, (e) #g, (f) #h, (g) #i, (h) #j, (i) #k.

Close modal

Comparison of the results of energy dissipation rate and water surface superelevation coefficient

Figure 15 shows the comparison of the energy dissipation rate and the water surface superelevation coefficient in Run 5 and Run OPC. In Run OPC, the energy dissipation rate (0.591) increased by 18.67%, and the water surface superelevation coefficient (0.306) decreased by 26.14% compared with those in Run 5. Thus, the energy dissipation rate and water surface elevation coefficient had better performance in Run OPC than in Run 5.
Figure 15

Comparisons between Run OPC and Run 5: (a) energy dissipation rate, (b) water surface superelevation coefficient.

Figure 15

Comparisons between Run OPC and Run 5: (a) energy dissipation rate, (b) water surface superelevation coefficient.

Close modal

The above numerical results show that R-SEDs' arrangement in the spillway bend can facilitate good energy dissipation and flow diversion. All the hydraulic indexes in Run OPC were superior to those in Run 5. These verified the reliability of the established multi-criteria evaluation system (i.e., the advantages of the recommended parameter combination) and the necessity of arranging R-SEDs in the bend.

This study focused on the R-SEDs arranged at the bend bottom of the curved spillway. A multi-criteria evaluation system was established for the comprehensive energy dissipation and flow diversion effects of R-SEDs. Based on hydrodynamic theory, orthogonal tests and numerical simulation, the six factors (average R-SED height h, R-SED spacing s, R-SED angle α, bend width w, bend centerline radius R and discharge flow rate Q) affecting the energy dissipation and diversion effects of the R-SEDs were analyzed to obtain an optimal parameter combination. The main conclusions are drawn as follows:

  • (i)

    Bend width (w) and bend centerline radius (R) are the significant factors affecting the energy dissipation effect of R-SEDs. These two factors are positively correlated with the energy dissipation rate. The factors are ranked in descending order in terms of their effects on R-SEDs' energy dissipation performance: bend width (w) > bend centerline radius (R) > R-SED angle (α) > R-SED spacing (s) > discharge flow rate (Q) > average R-SED height (h).

  • (ii)

    Average R-SED height (h), R-SED spacing (s) and bend centerline radius (R) are the significant factors affecting the flow diversion effect of R-SEDs. Average R-SED height (h) is negatively correlated with the water surface superelevation coefficient, while R-SED spacing (s) and bend centerline radius (R) are positively correlated. The factors are ranked in descending order in terms of their effects on R-SEDs' flow diversion performance: average R-SED height (h) > bend centerline radius (R) > R-SED spacing (s) > R-SED angle (α) > bend width (w) > discharge flow rate (Q).

  • (iii)

    Bend width (w), average R-SED height (h) and bend centerline radius (R) are the significant factors influencing the comprehensive energy dissipation and flow diversion effects. These three factors are positively correlated with relative closeness. The factors are ranked in descending order in terms of their effects on R-SEDs' comprehensive performance: bend width (w) > average R-SED height (h) > bend centerline radius (R) > R-SED angle (α) > R-SED spacing (s) > discharge flow rate (Q).

  • (iv)

    R-SEDs are simple in shape and convenient in construction. They can effectively improve the various undesirable hydraulic phenomena in the bend flow. The influence weights of the energy dissipation rate (0.629) and the water surface superelevation coefficient (0.371) indicate that the energy dissipation effect of the R-SEDs on the bend flow is more significant than the flow diversion effect.

  • (v)

    Among the 18 groups of orthogonal tests, Run 5 shows the highest overall rating. The run with the recommended parameter combination (h = 1.8 cm, s = 24 cm, α = 18°, R = 200 cm, w = 80 cm and Q = 20.0 L/s) from the multi-criteria evaluation system exhibits better hydraulic evaluation indexes than Run 5. For example, the energy dissipation rate (0.591) is 18.67% higher than that in Run 5. The water surface superelevation coefficient (0.306) is 26.14% lower than that in Run 5. The recommended parameter combination can provide a reference for the R-SED design of similar curved spillways based on actual project topography and geological conditions.

  • (vi)

    This study mainly considers the effects of the six key factors (average R-SED height h, R-SED angle α, R-SED spacing s, bend width w, bend centerline radius R and discharge flow rate Q) on the energy dissipation and flow diversion effects of R-SEDs. Further studies on other factors (such as bend angle and bottom slope) will be conducted subsequently. Regarding the R-SED design of the same type of curved spillways, the effects of relevant parameters on R-SEDs' energy dissipation and flow diversion effects in this study can provide a reference for selecting the parameter values. For example, based on the recommended parameter values, the main effect analysis results and engineering characteristics, relevant simulation tests can be performed to select the values of significant factors. For the secondary factors, combined with the recommended parameter values and the main effect analysis results, their values can be directly taken based on engineering experience or economic practicality. The relevant parameters used in this study were the dimensions of the physical laboratory model. Thus, these parameters should be scaled up according to the geometric scale (1:50) for actual engineering design.

The authors gratefully acknowledge the financial support from the National Natural Science Foundation of China (Grant numbers 52269019 and 51769037) and the University Research Program Innovation Team Project of Xinjiang Uygur Autonomous Region (Grant number XJEDU2017T004). We would also like to thank XSG Editing for their professional editing service.

This work was supported by the National Natural Science Foundation of China (Grant numbers 52269019 and 51769037) and University Research Program Innovation Team Project of Xinjiang Uygur Autonomous Region (Grant number XJEDU2017T004). The first author has received research support from Xinjiang Agricultural University.

All relevant data are included in the paper or its Supplementary Information.

The authors declare there is no conflict.

Anwar
H. O.
1986
Turbulent structure in a river bend
.
Journal of Hydraulic Engineering
112
(
8
),
657
669
.
https://doi.org/10.1061/(asce)0733-9429(1986)112:8(657)
.
Bathurst
J. C.
,
Hey
R. D.
&
Thorne
C. R.
1979
Secondary flow and shear stress at river bends
.
Journal of the Hydraulics Division, ASCE
105
,
1
64
.
https://doi.org/10.1061/jyceaj.0005285
.
Booij
R.
2003
Measurements and large eddy simulations of the flows in some curved flumes
.
Journal of Turbulence
4
(
1
),
008
.
https://doi.org/10.1088/1468-5248/4/1/008
.
Damarnegara
S.
,
Wardoyo
W.
,
Perkins
R.
&
Vincens
E.
2020
Computational fluid dynamics (CFD) simulation on the hydraulics of a spillway
.
IOP Conference Series: Earth and Environmental Science
437
(
1
),
012007
.
https://doi.org/10.1088/1755-1315/437/1/012007
.
De Vriend
H. J.
&
Geldof
H. J.
1983
Main flow velocity in short river bends
.
Journal of Hydraulic Engineering
109
(
7
),
991
1011
.
https://doi.org/10.1061/(ASCE)0733-9429(1983)109:7(991)
.
Dietrich
W. E.
,
Smith
J. D.
&
Dunne
T.
1979
Flow and sediment transport in a sand bedded meander
.
The Journal of Geology
87
(
3
),
305
315
.
https://doi.org/10.1086/628419
.
Engel
F. L.
&
Rhoads
B. L.
2016
Three-dimensional flow structure and patterns of bed shear stress in an evolving compound meander bend
.
Earth Surface Processes and Landforms
41
(
9
),
1211
1226
.
https://doi.org/10.1002/esp.3895
.
Ghazanfari-Hashemi
R. S.
,
Montazeri Namin
M.
,
Ghaeini-Hessaroeyeh
M.
&
Fadaei-Kermani
E.
2019
A numerical study on three-dimensionality and turbulence in supercritical bend flow
.
International Journal of Civil Engineering
18
(
3
),
381
391
.
https://doi.org/10.1007/s40999-019-00471-w
.
Gu
L.
,
Zhang
S.
,
He
L.
,
Chen
D.
,
Blanckaert
K.
,
Ottevanger
W.
&
Zhang
Y.
2016
Modeling flow pattern and evolution of meandering channels with a nonlinear model
.
Water
8
(
10
),
418
.
https://doi.org/10.3390/w8100418
.
Han
S. S.
,
Ramamurthy
A. S.
&
Biron
P. M.
2011
Characteristics of flow around open channel 90 bends with vanes
.
Journal of Irrigation and Drainage Engineering
137
(
10
),
668
676
.
https://doi.org/10.1061/(asce)ir.1943-4774.0000337
.
Hu
C.
,
Yu
M.
,
Wei
H.
&
Liu
C.
2019
The mechanisms of energy transformation in sharp open-channel bends: analysis based on experiments in a laboratory flume
.
Journal of Hydrology
571
,
723
739
.
https://doi.org/10.1016/j.jhydrol.2019.01.074
.
Huai
W.
,
Li
C.
,
Zeng
Y.
,
Qian
Z.
&
Yang
Z.
2012
Curved open channel flow on vegetation roughened inner bank
.
Journal of Hydrodynamics
24
(
1
),
124
129
.
https://doi.org/10.1016/s1001-6058(11)60226-6
.
Jin
Y.
&
Steffler
P. M.
1993
Predicting flow in curved open channels by depth-averaged method
.
Journal of Hydraulic Engineering
119
(
1
),
109
124
.
https://doi.org/10.1061/(asce)0733-9429(1993)119:1(109)
.
Johannesson
H.
&
Parker
G.
1989
Velocity redistribution in meandering rivers
.
Journal of Hydraulic Engineering
115
(
8
),
1019
1039
.
https://doi.org/10.1061/(asce)0733-9429(1989)115:8(1019)
.
Kells
J. A.
&
Smith
C. D.
1991
Reduction of cavitation on spillways by induced air entrainment
.
Canadian Journal of Civil Engineering
18
(
3
),
358
377
.
https://doi.org/10.1139/l91-047
.
Kim
J. S.
,
Baek
D.
&
Park
I.
2020
Evaluating the impact of turbulence closure models on solute transport simulations in meandering open channels
.
Applied Sciences
10
(
8
),
2769
.
https://doi.org/10.3390/app10082769
.
Li
Z.
,
Liu
Z.
,
Wang
H.
,
Chen
Y.
,
Li
L.
,
Wang
Z.
&
Zhang
D.
2022
Investigation of aerator flow pressure fluctuation using detached eddy simulation with VOF method
.
Journal of Hydraulic Engineering
148
(
1
),
04021052
.
https://doi.org/10.1061/(ASCE)HY.1943-7900.0001953
.
Maatooq
J. S.
&
Hameed
L. K.
2020
2D model to investigate the morphological and hydraulic changes of meanders
.
Engineering and Technology Journal
38
(
1
),
9
19
.
https://doi.org/10.30684/etj.v38i1A.95
.
Martin-Vide
J. P.
,
Roca
M.
&
Alvarado-Ancieta
C. A.
2010
Bend scour protection using riprap
.
Proceedings of the Institution of Civil Engineers-Water Management
163
(
10
),
489
497
.
https://doi.org/10.1680/wama.2010.163.10.489
.
Ministry of Water Resources of China
1995
Test Regulation for Special Hydraulic Model, SL 156 ∼ 165-95
.
China Water & Power Press
,
Beijing
.
(In Chinese).
Ministry of Water Resources of China
2012
Specification for Normal Hydraulic Model Test, SL 155-2012
.
China Water & Power Press
,
Beijing
.
(In Chinese).
Ministry of Water Resources of China
2016
Calibration Method of Common Instruments for Hydraulic and River Model Test, SL 233-2016
.
China Water & Power Press
,
Beijing
.
(In Chinese).
Molls
T.
&
Chaudhry
M. H.
1995
Depth-averaged open-channel flow model
.
Journal of Hydraulic Engineering
121
(
6
),
453
465
.
https://doi.org/10.1061/(asce)0733-9429(1995)121:6(453)
.
Moncho-Esteve
I. J.
,
García-Villalba
M.
,
Muto
Y.
,
Shiono
K.
&
Palau-Salvador
G.
2018
A numerical study of the complex flow structure in a compound meandering channel
.
Advances in Water Resources
116
,
95
116
.
https://doi.org/10.1016/j.advwatres.2018.03.013
.
Odgaard
A. J.
&
Bergs
M. A.
1988
Flow processes in a curved alluvial channel
.
Water Resources Research
24
(
1
),
45
56
.
https://doi.org/10.1029/wr024i001p00045
.
Pazooki
P.
,
Hamedi
A.
,
Torabi
M.
,
Zeidi
S.
&
Vosoughifar
H.
2020
Effect of changing the height of final step of the stepped chute on the flow profile in stilling basin using the VOF method
.
Applied Water Science
10
(
7
),
1
9
.
https://doi.org/10.1007/s13201-020-01262-8
.
Pradhan
A.
,
Kumar Khatua
K.
&
Sankalp
S.
2018
Variation of velocity distribution in rough meandering channels
.
Advances in Civil Engineering
2018
,
1
12
.
https://doi.org/10.1155/2018/1569271
.
Qin
C.
,
Shao
X.
&
Zhou
G.
2016
Comparison of two different secondary flow correction models for depth-averaged flow simulation of meandering channels
.
Procedia Engineering
154
,
412
419
.
https://doi.org/10.1016/j.proeng.2016.07.507
.
Ramamurthy
A. S.
,
Han
S. S.
&
Biron
P. M.
2013
Three-dimensional simulation parameters for 90 open channel bend flows
.
Journal of Computing in Civil Engineering
27
(
3
),
282
291
.
https://doi.org/10.1061/(asce)cp.1943-5487.0000209
.
Ranjan
R.
,
Ahmad
Z.
&
Asawa
G. L.
2006
Effect of spacing of submerged vanes on bed scour around river bends
.
ISH Journal of Hydraulic Engineering
12
(
2
),
49
65
.
https://doi.org/10.1080/09715010.2006.10514831
.
Schreiner
H. K.
,
Rennie
C. D.
&
Mohammadian
A.
2018
Trajectory of a jet in crossflow in a channel bend
.
Environmental Fluid Mechanics
18
(
6
),
1301
1319
.
https://doi.org/10.1007/s10652-018-9594-8
.
Seo
I. W.
&
Shin
J.
2018
Two-dimensional modeling of flow and contaminant transport in meandering channels
.
EPiC Series in Engineering
3
,
1911
1918
.
https://doi.org/10.1201/b10999-55
.
Seyedashraf
O.
&
Akhtari
A. A.
2015
Flow separation control in open-channel bends
.
Journal of the Chinese Institute of Engineers
39
(
1
),
40
48
.
https://doi.org/10.1080/02533839.2015.1066942
.
Shaheed
R.
,
Mohammadian
A.
&
Yan
X.
2021
A review of numerical simulations of secondary flows in river bends
.
Water
13
(
7
),
884
.
https://doi.org/10.3390/w13070884
.
Thomson
J.
1876
On the origin of windings of rivers in alluvial plains, with remarks on the flow of water round bends in pipes
.
Proceedings of the Royal Society of London
25
(
171–178
),
5
8
.
https://doi.org/10.1098/rspl.1876.0004
.
Vaghefi
M.
,
Akbari
M.
&
Fiouz
A. R.
2015
An experimental study of mean and turbulent flow in a 180 degree sharp open channel bend: secondary flow and bed shear stress
.
KSCE Journal of Civil Engineering
20
(
4
),
1582
1593
.
https://doi.org/10.1007/s12205-015-1560-0
.
Wang
H.
,
Wang
X.
,
Bi
L.
,
Wang
Y.
,
Fan
J.
,
Zhang
F.
,
Hou
X.
,
Cheng
M.
,
Hu
W.
,
Wu
L.
&
Xiang
Y.
2019
Multi-objective optimization of water and fertilizer management for potato production in sandy areas of northern China based on TOPSIS
.
Field Crops Research
240
,
55
68
.
https://doi.org/10.1016/j.fcr.2019.06.005
.
Wang
L.
,
Zhang
X.
,
Wan
R.
,
Xu
Q.
&
Qi
G.
2022
Optimization of the hydrodynamic performance of a double-vane otter board based on orthogonal experiments
.
Journal of Marine Science and Engineering
10
(
9
),
1177
.
https://doi.org/10.3390/jmse10091177
.
Yan
X.
,
Rennie
C. D.
&
Mohammadian
A.
2020
A three-dimensional numerical study of flow characteristics in strongly curved channel bends with different side slopes
.
Environmental Fluid Mechanics
20
(
6
),
1491
1510
.
https://doi.org/10.1007/s10652-020-09751-9
.
Yang
J.
,
Zhang
J.
,
Zhang
Q.
,
Teng
X.
,
Chen
W.
&
Li
X.
2019
Experimental research on the maximum backwater height in front of a permeable spur dike in the bend of a spillway chute
.
Water Supply
19
(
6
),
1841
1850
.
https://doi.org/10.2166/ws.2019.061
.
Ye
J.
&
McCorquodale
J. A.
1998
Simulation of curved open channel flows by 3d hydrodynamic model
.
Journal of Hydraulic Engineering
124
(
7
),
687
698
.
https://doi.org/10.1061/(asce)0733-9429(1998)124:7(687)
.
Yuan
Y.
,
Jin
R.
,
Tang
L.
&
Lin
Y.
2022
Optimization design for the centrifugal pump under non-uniform elbow inflow based on orthogonal test and GA_PSO
.
Processes
10
(
7
),
1254
.
https://doi.org/10.3390/pr10071254
.
Zhang
Q.
,
Diao
Y.
,
Zhai
X.
&
Li
S.
2015
Experimental study on improvement effect of guide wall to water flow in bend of spillway chute
.
Water Science and Technology
73
(
3
),
669
678
.
https://doi.org/10.2166/wst.2015.523
.
Zhang
H.
,
Mu
Z.
,
Fan
F.
&
Li
F.
2022a
Analysis of effects of rough strip energy dissipator on hydraulic property of bend flow
.
Geofluids
2022
.
https://doi.org/10.1155/2022/4260540
.
Zhang
H.
,
Mu
Z.
,
Fan
F.
&
Li
F.
2022b
Analysis of multifactor influence model on energy dissipation rate of rough-strips energy dissipator in the spillway bend
.
Advances in Materials Science & Engineering
2022
.
https://doi.org/10.1155/2022/2017448
.
Zhou
J.
,
Shao
X.
,
Wang
H.
&
Jia
D.
2017
Assessment of the predictive capability of RANS models in simulating meandering open channel flows
.
Journal of Hydrodynamics, Series B
29
(
1
),
40
51
.
https://doi.org/10.1016/s1001-6058(16)60714-x
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the original work is properly cited (http://creativecommons.org/licenses/by-nc-nd/4.0/).